You are on page 1of 43

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233162245

Evaporative Cooling of Water in a Small Vessel Under Varying Ambient Humidity

Article  in  International Journal of Green Energy · October 2006


DOI: 10.1080/01971520600704654

CITATIONS READS
24 4,390

4 authors, including:

Ashutosh Mittal Siddharth Chatterjee


National Renewable Energy Laboratory State University of New York College of Environmental Science and Forestry
52 PUBLICATIONS   917 CITATIONS    57 PUBLICATIONS   479 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Production of biogas from co-digestion of paper waste and manure View project

Biomass conversion View project

All content following this page was uploaded by Siddharth Chatterjee on 27 May 2014.

The user has requested enhancement of the downloaded file.


Evaporative Cooling of Water in a Small Vessel under
Varying Ambient Humidity
Ashutosh Mittala, Tarun Katariaa,1 , Gautam K. Dasb, and Siddharth G. Chatterjeea, 2
a
Faculty of Paper Science and Engineering, SUNY College of Environmental Science and
Forestry, 1 Forestry Drive, Syracuse, New York 13210, U.S.A.
b
118/A Acharya J. C. Bose Road, Kolkata 700014, India

ABSTRACT

Evaporative cooling of water in a small porous clay vessel was studied under

controlled humidity conditions. In steady-state experiments performed at an ambient

temperature of 23 °C, the cooling effect increased from 4.7 to 8.3 °C as the ambient

relative humidity decreased from 60 to 15%. External heat and mass transfer coefficients,

estimated from the steady-state measurements, were used in mathematical models to

predict the experimentally observed transient temperature variation of the water under

ramp changes of the ambient relative humidity. With a prototypical cool chamber

containing water tested in Kolkata, India under an ambient temperature of 34.5–35 °C,

the cooling effect reached a maximum of 7 °C between 3 and 3:30 PM and then declined

to 4.5 °C around 6 PM.

Keywords: Evaporative cooling; Heat transfer; Mass transfer; Porous media; Cool

chamber

1
Presently at Portellus, Inc., 2010 Main St., Suite 450, Irvine, CA 92614, U.S.A.
2
Correspondence: S. G. Chatterjee, Faculty of Paper Science and Engineering, SUNY College of
Environmental Science and Forestry, 1 Forestry Drive, Syracuse, NY 13210, U.S.A., E-mail:
schatterjee@esf.edu

1
INTRODUCTION

Evaporative cooling, a technique that has been in use for several centuries in the

Middle East and Africa, is a passive cooling process in which a body or an object is

cooled by the evaporation of water from its surface. Since only ambient energy is utilized

for the evaporation, evaporative cooling is a very promising method that can be used in

the design of storage facilities for perishable food products like fruits and vegetables and

for air conditioning in rural areas of developing countries, especially when the local

weather conditions are hot and dry. This cooling technique should be especially relevant

at present when depleting fossil fuel reserves and environmental problems caused by the

rapid use of such fuels are of mounting global concern. Since evaporative cooling has

diverse applications, we present a brief review of some of the recent literature in this

area.

Aimiuwu (1992, 1993) found the long-term temperature of water in a porous

ceramic pot to be 10.4–15 °C below the ambient temperature and to have a smaller daily

variation than the external temperature. Taha et al. (1994) observed a temperature

depression of 10–13 °C in a box shaped evaporatively cooled chamber (ECC) constructed

of zinc whose outer surface was covered by continuously wetted charcoal layers. Fruits

and vegetables like pomegranates, bananas, mangoes, tomatoes, and potatoes not only

had significantly increased shelf lives (by a factor of 1.2–2.7) but also exhibited good

appearance and physiological properties when stored in an ECC compared to storage

under ambient conditions (Waskar et al., 1999; Dzivama et al., 1999; Kumar et al., 1999;

Uppal, 1999; Thakur et al., 2002; Dhemre and Waskar, 2003; Mordi and Olorunda,

2003). Anyanwu (2004) measured the transient response of an ECC to changes in the

2
ambient relative humidity (RH) and temperature during dry and wet seasons. The box

shaped ECC had two clay walls with continuously wetted coconut fibre filling the gap

between the walls. On average, the temperature inside the ECC was 1–8.2 °C and 3–12

°C lower than the ambient value during the wet and dry seasons, respectively. The ECC

also increased the shelf lives of tomatoes and pumpkins by factors of 2.9 and 5 above

their open-air storage values. Dash and Chandra (2001) developed a mathematical model

to study the influence of structural and operational parameters on the interior

environment of an ECC, which was mainly affected by the rate of evaporation on the

outer surfaces of the ECC and the infiltration/exfiltration rates through the chamber.

Upchurch and Mahan (1988) found the characteristic leaf temperature of well-

watered cotton plants to be 27 ± 2 °C as the ambient air temperature varied from 27–40

°C. They attributed this homeothermic behavior in a eurythermal environment to

evaporative cooling by transpiring leaf surfaces and active water use by the plant through

which it is able to control its temperature variation. According to Prange (1996), insects

like bees, grasshoppers, beetles, and cockroaches use evaporative cooling as a

thermoregulatory mechanism under certain conditions to avoid heat stress in hot

environments by which they maintain a body temperature below 48 °C.

Because of its potential for significant savings in energy consumption,

evaporative cooling is drawing increasing attention in indoor air conditioning. From a

numerical simulation, Fu et al. (1990) concluded that the temperature of a structure

subjected to a severe thermal radiative heat flux decreased due to the presence of an

evaporating water film on the structure’s surface. Kassem (1994) studied theoretically

and experimentally a farm structure in which air, humidified and cooled via passing

3
through a vertically oriented wetted pad, entered at one end of the building and exited

from the opposite end. His calculations revealed that increasing the water evaporation

efficiency of the pad and regulating the air and water intake rates to follow the diurnal

variation of the ambient temperature would lead to substantial savings in the energy and

water consumptions. Giabaklou and Ballinger (1996) and Ghiabaklou (2003) proposed a

system that used the evaporative cooling effect of water falling vertically along filaments

or guides to cool and humidify the air entering a building. For hot and semi-arid regions

like Wagga Wagga, New South Wales, Australia, and Teheran, Iran, their model

predicted indoor average maximum temperatures to be 9.9 °C and 11.7 °C lower than the

temperature of the incoming ambient air, respectively. Their simulation model also

indicated that the evaporative cooling system would result in a more stable indoor

temperature and thermal comfort level as estimated from the predicted mean vote. In an

experimental investigation of porous ceramic evaporators for building cooling, Ibrahim et

al. (2003) measured dry bulb temperature drops of 6–8 °C with a 30% increase in the RH

of the inlet air. The cooling effect was enhanced by a high porosity of the evaporator,

increased water supply pressure, and a single (as compared with a twin) row of

evaporators in the air duct. Hollow fibre membrane contactors, which have a large mass

transfer area per unit volume, with water and air flowing inside and outside the

microporus tubes, respectively, also have a good potential for space air conditioning and

have been investigated by Bergero and Chiari (2001) and Johnson et al. (2003). Dai and

Sumathy (2002) have presented a theoretical model in which wet honeycomb paper is

used as the packing material through which the air stream to be cooled and humidified

flows in a cross-flow fashion. Tang and Etzion (2004) and Cheikh and Bouchair (2004)

4
have developed dynamic models for predicting the thermal performance of roof ponds of

special designs for cooling buildings in hot and arid climates. From experimentation with

a variety of test structures, Nahar et al. (1999) recommended a roof with a shallow pond

and movable insulation for buildings in rural arid areas.

Zalewski and Gryglaszewski (1997) developed a heat and mass transfer model for

evaporative fluid coolers in which water was sprayed over the surface of the tubes while

air flowed outside the tubes in a countercurrent fashion. A correction to the mass transfer

coefficient (initially calculated from the Lewis relation) was introduced which made

model predictions for the thermal performance of the cooler agree quite closely with

experimental measurements. From an extensive experimental investigation, Armbruster

and Mitrovic (1998) developed correlations for the Nusselt number and temperature

decrease (due to evaporative cooling) for water falling freely from a horizontal tube to the

next one below it. They observed that the major part of the cooling effect occurred during

the free fall of the water between adjacent tubes rather than when it flowed as a film

around the tube. From a theoretical analysis of the evaporation of water flowing in a thin

film around the fin of an air-cooled heat exchanger, Song et al. (2003) concluded that

evaporation would significantly increase the cooling effect.

Chuntranuluck et al. (1998a) developed an algebraic method for predicting the

chilling time, an important consideration in the food industry, for a regular-shaped food

(infinite slab, infinite cylinder and sphere), which cools by convection and evaporation at

its surface. Their model agreed fairly well with experimental measurements of the

chilling time of a food analogue (Chuntranuluck et al., 1998b) and (peeled and unpeeled)

carrots (Chuntranuluck et al., 1998c). A more elegant model, which describes the cooling

5
of a high-moisture cylindrically shaped food via convection and evaporation at its

surface, has recently been proposed by van der Sman (2003). This model was based on

numerical calculations with the heat diffusion equation inside the food body, which

showed that except for a short initial time period, the average temperature remained in a

fixed location. However, van der Sman did not present any comparison of the predictions

of his model with experimental data.

In the present study, the following tasks are accomplished: (1) We test van der

Sman’s model against experimental measurements of the temperature of water contained

in a small porous clay pot kept inside a controlled humidity chamber whose RH was

changed from one prescribed level to another in a ramped fashion. The mass and heat

transfer coefficients outside the vessel, which are required in the model, were estimated

from steady-state experiments. (2) We also develop a very simple mathematical model

that describes the temperature evolution of an object cooled by convection and

evaporation at its surface under a dynamically changing ambient RH. Unlike the model of

van der Sman, this model requires the ratio of the mass and heat transfer coefficients and

has a closed form solution.

Good agreement between the theoretical predictions and experimental data

indicate the usefulness of the models for the rational design of evaporatively cooled

chambers and hollow fiber membrane or porous ceramic units for room air conditioning.

Finally, we present some results of the evaporative cooling effect obtained with a

prototypical ECC under tropical conditions.

6
VAN DER SMAN’S MODEL

The chief assumptions in van der Sman’s model (van der Sman, 2003) which

describes the cooling of a regular-shaped body are: (1) heat transport inside the body

occurs only by conduction along the radial direction x (defined as the direction outwardly

normal from the axis of symmetry of the body), (2) the body is water logged with

evaporation occurring at its outer surface, and (3) thermal properties of the body are

independent of temperature. The cooling of the body as a function of time t is then

described by (van der Sman, 2003)

dTavg
ρ eff c p ,eff V = −φ ext (1)
dt

where

∂T
φext = − keff Aext = hext Aext (Ts − Tamb ) + λ (Ts ) k f ,ext Aext (cs (Ts ) − camb ) (2)
∂x x=ap

Note that in Eq. (2) it has been assumed that the outer surface of the body is completely

wet, i.e., water activity is unity. The assumption of a porous medium completely

saturated with water significantly simplifies the theoretical treatment of the evaporative

cooling process. (A model for moisture evaporation and migration in a thin unsaturated

porous packed bed has been presented by Liu et al., 1997.) From the ideal gas law, it can

be shown that

M w P sat (Tamb )  P sat (Ts ) Tamb RH amb 


c s (Ts ) − c amb =  −  (3)
R gas Tamb  P sat (Tamb ) Ts 100 

According to the model of van der Sman (2003)

φ ext = hint Aext (Tavg − Ts ) (4)

where hint is an internal heat transfer coefficient defined by

7
k eff
hint = (5)
dc

Here, dc is a characteristic length that depends on the geometry of the body (slab, cylinder

or sphere) and is given in Table 1 as a function of the half thickness or radius of the body,

ap. From Eqs. (2) and (4) we can derive the relation between the average and surface

temperatures of the body to be

hext λ (T )k
Tavg = Ts + (Ts − Tamb ) + s f ,ext (c s (Ts ) − camb ) (6)
hint hint

and, therefore,

 dλ (Ts )
(cs (Ts ) − camb ) + λ (Ts ) dc s (Ts ) 
dTavg hext k f ,ext 
= f (Ts ) = 1 + +  (7)
dTs hint hint  dTs dTs 

From Eqs. (1) and (2) it may be easily shown that the variation of the surface temperature

with time is described by

dTs g (Ts )
= (8)
dt f (Ts )

where

g (Ts ) =
Aext
[
ρ eff c p ,eff V
]
hext (Tamb − Ts ) − λ (Ts )k f ,ext (c s (Ts ) − c amb ) (9)

For moderate temperature ranges, the vapor pressure of water Psat and the latent heat of

vaporization λ can be well represented by

P sat (T ) = c1e c2T (10)

λ (T ) = c3 + c 4T (11)

8
where c1–c4 are constants. In this case, the functions f and g are given by

c T
hext k f , ext M wc1e 2 amb   ec 2Ts Tamb RH amb  (c3 + c4Ts )Tamb ec 2Ts  c2Ts − 1 
f (Ts ) = 1 + + c4  c 2Tamb − +  
hint hint RgasTamb   e Ts 100  e c 2Tamb  Ts  
2

(12)

 (c3 + c4Ts )k f ,ext M wc1ec2Tamb  ec 2Ts Tamb RH amb 


g (Ts ) = ( )
Aext
 ext amb
h T − T −  c T −  (13)
ρ eff c p , eff V
s
 RgasTamb  e T 100 
2 amb
s

The initial condition for Eq. (8) is

Ts = Tavg = Tini at t = 0 (14)

Note that the ambient temperature Tamb and relative humidity RHamb as well as the

external heat and mass transfer coefficients hext and kf,ext can be arbitrary functions of

time t.

Utilizing Eqs. (2), (4) and (5), Eq. (1) can be recast as

= hi (Ts − Tavg )
dTavg
(15)
dt

where hi is an effective internal heat transfer coefficient defined by

 keff  Aext A
hi =   = α eff ext (16)
ρ c
 eff p ,eff  Vd c Vd c

Equation (15) is the well-known linear driving force (LDF) approximation that represents

the heat transfer process inside the body. The LDF concept has been used previously to

describe intraparticle diffusion of adsorbate in porous adsorbent pellets; it is also referred

to as the Glueckauf approximation (Tien, 1994). Note that the characteristic length dc in

the Glueckauf approximation is different from that in van der Sman’s model; expressions

for dc for both approximations are given in Table 1 for different geometries (slab,

cylinder and sphere). The Glueckauf approximations for the slab and cylinder can be

9
derived by assuming a quadratic temperature profile inside the body and minimizing the

weighted square of the residual similar to the method presented by Tien (1994) for the

case of a sphere. This mathematical procedure, however, does not lead to any insight into

the physical nature of the LDF approximation unlike the explanation provided by van der

Sman (2003).

Under steady-state or equilibrium conditions, Tavg = Ts = Teq, and Eq. (6) becomes

λ (Teq ) k f ,ext
Tamb − Teq =
hext
(c (T ) − c )
s eq amb (17)

where Teq is the wet-bulb temperature (Treybal, 1980). The mass transfer coefficient kf,ext

is related to the steady-state weight loss rate of the body Wloss (due to water evaporation)

by

Wloss = k f ,ext Aext ( cs − camb ) (18)

Once kf,ext has been determined from Eq. (18), the heat transfer coefficient hext can be

obtained from Eq. (17), provided values of Tamb, Teq and RHamb are known.

A photograph and schematic diagram of the small pot, made of highly porous red

clay, which was used in the controlled humidity experiments, are shown in Fig. 1. The

pot, which had cylindrical symmetry, consisted of an upper cylindrical portion and a

lower frustum. The area of its external surface (excluding the top and the base) can be

calculated from

Aext = 2π rcyl lcyl + π ( rtop + rbot ) h f 2 + ( rtop − rbot ) + π (rcyl 2 − rtop 2 )


2
(19)

The clay pot was conceptually modeled as a regular cylinder having an equivalent radius

given by

10
Aext
Requiv = (20)
2π (lcyl + h f )
For a cylinder, the characteristic length dc, which is related to the internal heat transfer

coefficient hint by Eq. (5), is equal to Requiv/3 in the van der Sman model (Table 1). If we

neglect the effect of curvature, which will be valid for a thin wall (for the clay pot used in

our experiments, twall/Requiv = 0.14), the effective thermal conductivity keff for use in Eq.

(5) can be estimated from

Requiv
keff = (21)
t wall Rwater
+
kclay k water

where

Rwater = Requiv − t wall (22)

For the clay pot containing water used in our work, the dominant resistance to internal

heat transport was offered by the water, hence, the value of keff was close to that of kwater.

The quantity ρ eff c p ,eff V in Eq. (1) for our system of the clay pot containing water

can be expressed as

ρ eff c p ,eff V = mclay c p ,clay + mwater c p ,water (23)

Although the mass of water in the pot mwater changed continuously during an experiment

in which the ambient relative humidity RHamb was varied in a transient fashion, in our

calculations we have used an average value of mwater since the amounts of water

evaporated were only 4.9 and 6.5% in the two dynamic experimental runs reported in this

paper.

11
In our work in which Tamb was kept constant, we found from steady-state

experiments that the heat and mass transfer coefficients could be adequately represented

by the expressions

k f ,ext = c5 RH amb + c6 (24)

k f ,ext
= c7 RH amb + c8 (25)
hext

where c5–c8 are constants. In the dynamic experiments, the ambient relative humidity

RHamb was varied linearly from an initial equilibrium level of RHi at time t = 0 down to a

final value of RHf at a rate r. The time tr of the duration of the ramp is then given by

RH i − RH f
tr = (26)
r

For t > tr, RHamb was maintained at a level equal to RHf. Therefore, RHamb can be

described by

 RH i − RH f 
RH amb (t ) = RH i −  t for 0 ≤ t ≤ tr (27a)
 tr 

= RHf for t > tr (27b)

Equation (8), subject to the initial condition expressed by Eq. (14), can be

integrated numerically in order to obtain the surface temperature Ts of the body as a

function of time t after which the average temperature of the body Tavg may be calculated

from Eq. (6).

In the steady-state and dynamic experiments to be described later, a

thermocouple, located approximately midway between the top and bottom of the clay pot

along its central axis of symmetry, was used to measure the water temperature, Tcntr.

12
Thus, we need a relation between the average temperature of the water Tavg, predicted

from van der Sman’s model described above, and Tcntr. Let us assume a power-law

variation for the temperature of the water in the pot given by

Tw ( x, t ) = a (t ) + b (t ) x n (28)

where a and b are functions of time and n (≥ 1) is a constant. Noting that for a cylinder

having a radius ap (= Requiv)

ap
2
Tavg (t ) = 2 ∫ T ( x, t ) xdx
w (29)
ap 0

it may be easily shown that

2 (Tavg − Ts )
Tcntr = Tavg + (30)
n

∂T keff Aext (Tavg − Ts )


φext = − keff Aext = (31)
∂x x =a p a p / (n + 2)

where Tcntr and Ts are the values of Tw at x = 0 (central axis of cylinder) and at x = ap

(surface of cylinder), respectively.

It can be readily observed from Eqs. (4), (5), (31), and Table 1 that van der

Sman’s model results from a choice of n = 1. Although a value of n exactly equal to 1

does not satisfy the zero gradient condition of ∂Tw/∂x = 0 at x = 0, a value of n arbitrarily

close to 1 does satisfy this condition as can be seen from Eq. (28). Therefore, Eq. (30)

with a value of n = 1 was used to calculate the central temperature from the average and

surface temperatures, which were obtained from van der Sman’s model.

13
SIMPLE MODEL

In order to circumvent the theoretical and experimental complexity embodied in

the use of van der Sman’s model outlined above, we present a very simple model that

describes the variation of the average temperature of a body undergoing convective and

evaporative cooling at its surface under a dynamically varying ambient RH. Combining

Eqs. (1) and (4) gives

= − β (Tavg − Ts )
dTavg
(32)
dt

where the time constant β is given by

hint Aext
β= (33)
ρ eff c p ,eff V

We now assume that the surface of the body is in a state of quasi or pseudo

equilibrium with its surroundings so that its surface temperature can be calculated from

Eq. (17), i.e., Ts = Teq. This assumption, which actually contradicts the condition of

nonequilibrium at the body’s surface as expressed by Eq. (2), is similar to the assumption

of interfacial equilibrium usually made in the theoretical description of rate processes

(e.g., absorption of gases in liquids, adsorption of gases or liquids on solids, distillation,

etc). Any error resulting from this assumption can be absorbed into hint, which can be

considered to be an empirical factor. According to Eq. (17), Teq is a function of RHamb

and Tamb, which, in principal, may be arbitrary functions of time t. To use Eq. (17) in our

simple model, the inverse psychrometric ratio kf,ext/hext should be known unlike in van der

Sman’s model, which requires individual values of kf,ext and hext (which were measured in

this work). As mentioned earlier, kf,ext/hext can be estimated by measuring the steady-state

wet bulb temperature of the body under given ambient conditions while kf,ext, assuming an

14
empirical correlation is not available, has to be estimated by measuring the rate of weight

loss of the body due to evaporation, which is a much more difficult proposition for a

large object. Alternatively, the Lewis relation may be employed to estimate kf,ext/hext or

the air-water vapor psychrometric chart can be utilized to obtain Teq as a function of

RHamb and Tamb. In our work, the steady-state and dynamic experiments were conducted

in a room with a controlled climate (humidity and temperature) in which Tamb was kept

constant (22–23 °C). We found from the steady-state measurements that over a moderate

temperature range (14.7–18.7 °C in this work), the relation between Teq and RHamb could

be well described by

Teq = c9 RH amb + c10 (34)

where c9 and c10 are constants. Substituting Eqs. (27a) and (27b) into Eq. (34), we can

obtain the relation between Teq and t as

Teq = c11t + c12 for 0 ≤ t ≤ tr (35a)

= Teq,f for t > tr (35b)

where

c9 ( RH i − RH f )
c11 = − (36)
tr

c12 = c9 RH i + c10 (37)

and Teq,f is the (equilibrium) wet-bulb temperature corresponding to RHf [which can be

obtained by setting RHamb = RHf in Eq. (34)].

Using Eqs. (35a) and (35b), the solution of Eq. (32), subject to the initial

condition expressed by Eq. (14), can be easily shown to be given by

15
 c 
Tavg = c11t +  c12 − 11  (1 − e− β t ) + Tini e− β t for 0 ≤ t ≤ tr (38a)
 β 

( )
= Teq , f + Tavg ,tr − Teq , f e − β (t −tr ) for t > tr (38b)

where Tavg ,tr is the average temperature of the body at t = tr calculated by Eq. (38a).

Unlike in the case of van der Sman’s model, in the simple model it is not possible

to derive an expression relating Tcntr and Tavg due to the arbitrary assumption of

equilibrium at the body’s surface. Therefore, we postulate that the variation of Tcntr with

time has the same form as that expressed by Eqs. (38a) and (38b) in which hint [see Eq.

(33)] is considered to be an empirical factor that can be obtained by fitting the theoretical

Tcntr to its experimentally observed value. As justification for this assumption we note

that in the cooling of water contained in a flattened ellipsoidal ceramic porous pot,

Aimiuwu (1992) found the cooling curve to be well fitted by an expression of the type

given by Eq. (38b), and, as will be presented later, Eqs. (38a) and (38b) represent fairly

well the empirically observed transient behavior of Tcntr in the evaporative cooling of

water in the clay pot used in our experiments for ramp changes of the ambient RH.

EXPERIMENTAL PROCEDURES

A detailed description of the controlled humidity chamber (whose schematic

diagram is shown in Fig. 2) in which the steady-state and dynamic evaporative cooling

experiments were conducted (in Syracuse, New York) has been provided previously by

Chatterjee et al.(1997). As mentioned earlier, all experiments reported in this work were

performed at an ambient (i.e., humidity chamber) temperature Tamb of 22–23 °C. Inside

the humidity chamber, the clay pot (obtained from Syracuse Pottery, 6551 Pottery Road,

Warners, NY 13164-9756), almost filled with water, was placed on a piece of styrofoam

16
that was set on a pan suspended from a Mettler 200PM balance (accuracy: 1 mg), which

was connected to a computer. The top of the pot was sealed with a styrofoam lid and air

of a definite humidity, obtained by mixing wet and dry air in specific proportions,

continuously flowed around the pot. A Vaisala humidity sensor (accuracy: 1% RH)

monitored the chamber RH and temperature and sent its signal to the computer that

maintained the RH in the chamber at a prescribed level by controlling the wet and dry

airflows. A factory-calibrated thermocouple was inserted into the pot through the lid at its

center in order to measure the temperature Tcntr of the water inside. The thermocouple,

whose calibration was also tested in house, had an accuracy of 0.1 °C. The weight (loss)

of the cup and the RH of the chamber were monitored as a function of time by a

computerized data acquisition system.

The steady-state evaporative cooling experiments, which were performed in order

to obtain kf,ext and hext as a function of the chamber RH (i.e., RHamb), were conducted at

levels of 60, 45, 30, and 15% RH, respectively (see Table 4). At a prescribed value of

RHamb, the clay pot containing water was allowed to attain a state of equilibrium (i.e., no

temperature gradient in the water) with the ambient. This generally took 40-60 min and

equilibrium was assumed to be established when Tcntr attained a steady value (equal to

Teq). The slope of the steady-state (i.e., linear) part of the weight-loss curve was used to

estimate Wloss. Two experimental replicates were used at the prescribed level of RHamb

and average values of Wloss, Teq and Tamb of the two replicates were used to calculate the

ratio kf,ext/hext and kf,ext from Eqs. (17) and (18), respectively. With kf,ext and kf,ext/hext

known, hext could be calculated.

17
In the dynamic evaporative cooling experiments, the clay pot containing water

(suspended from the balance) was subjected to RH ramps inside the controlled humidity

chamber. Before the ramp was begun, RHamb was maintained at a value of RHi and the

pot was allowed to attain equilibrium. Thereafter, RHamb was decreased at a constant rate

(0.5 and 1% RH/min) until a final desired RH value, RHf, was attained in the chamber.

This final RH level was maintained for some more time so that the pot essentially

attained equilibrium. From the start of the RH ramp until the end of the experiment, the

temperature of the water in the pot Tcntr was recorded manually at periodic intervals

(generally every 5, 10 or 15 min). An average of the mass of water contained in the pot at

the beginning and at the end of the experiment was used as the value of mwater in Eq. (23).

In order to test the efficacy of evaporative cooling under tropical conditions, a

prototype of an ECC, shown in Fig. 3, was tested in Kolkata, India. This cool chamber

consisted of a copper vessel (containing water) wrapped in a piece of wet cloth that was

placed between two small fans. Periodically, a thermometer (accuracy: 0.1 °C) was used

to measure the temperature of the water Twater near the top and center of the vessel while

an analog Hucer hygrometer (accuracy: 1% RH) measured the ambient humidity in the

room during a summer afternoon of May 2004.

RESULTS AND DISCUSSION

Table 2 lists the various physical properties and relations while Table 3 reports

the parameters of the clay pot used in this work.

Table 4 and Figs. 4-6 exhibit results of the steady-state evaporative cooling

experiments. To obtain an idea of the errors involved in the experimental measurements,

consider the data reported in Table 4. The maximum differences in RHamb, Tamb and Teq

18
between the two replicates are 0.1% RH, 0.3 °C and 0.3 °C, respectively. The maximum

deviations of Wloss from its average values are 4.1% and 4.5% for the first and second

replicates, respectively. On average, the steady-state weight loss rate increased by about

19% as RHamb was reduced from 60 to 30%; it then decreased slightly with a further

reduction of RHamb to 15%. The following speculation can be advanced for this behavior:

As RHamb was reduced from 60%, the driving force for mass transfer (cs – camb) increased

while the mass transfer coefficient (kf,ext) decreased [see Eq. (18) and Fig. 5] as a result of

the changed air flow inside the humidity chamber. However, the increase in the former

more than compensated for a decrease in the latter, with a consequent increase in Wloss.

As RHamb was decreased below a certain level, the reduction in kf,ext and a possible

reduction in the mass transfer driving force due to a decrease in cs (which is an

exponential function of Ts) was responsible for the observed decrease in Wloss.

Figure 4 shows the experimental kf,ext/hext as a function of the steady-state RHamb

and the linear correlation given by Eq. (25). At the four experimental values of RHamb of

15, 30, 45, and 60%, the quantity ρair,amb kf,ext/hext takes values of 0.40, 0.45, 0.59, and

0.64 g K/J, respectively, which can be compared to the value of 1 g K/J given by the

Lewis relation for the air–water vapor system. Here, ρair,amb is the ambient air density

calculated by

ρ air ,amb =
(P
amb − P sat (Tamb )) M air
(39)
RgasTamb

Figure 5 shows that the experimental kf,ext decreased from 1.06 to 0.44 to cm/s as

the steady-state RHamb was reduced from 60 to 15% (due to different air flow conditions

in the humidity chamber) and that the linear correlation given by Eq. (24) is a fairly good

19
representation of the experimental data. The variation of the experimental steady-state

equilibrium temperature of the clay pot Teq is plotted in Fig. 6 as a function of the steady-

state RHamb along with the correlation given by Eq. (34). The experimental value of the

temperature drop ∆T = Tamb – Teq, which represents the cooling effect due to evaporation,

increased in the sequence 4.7, 6.3, 7.1, and 8.3 °C as RHamb was reduced from 60 to15%

in steps of 15% RH (Table 4). Thus, in contrast to the moderate effect of RHamb on Wloss

mentioned earlier, ∆T increased by about 77% as RHamb was lowered from 60 to 15%.

Figures 7 and 8 exhibit results of the dynamic evaporative cooling experiments in

which RHamb was changed from 60 to 15% linearly with time. The values of the ramp

rates and average ambient temperature Tamb in the two experimental runs shown in Figs. 7

and 8 were 0.5 and 1% RH/min and 22.3 and 22.2 °C, respectively. We note that no

fitting parameter was employed in the van der Sman model in order to match theoretical

predictions of the transient Tcntr with its experimental values. However, in the simple

model, a value of hint = 1.817 × 10-3 W/(cm2 K) was used to match theory with

experiment. Overall, both models capture the trend of the experimental variation of Tcntr

with time reasonably well, however, there is an over prediction of 1.4 and 1.2 °C by the

van der Sman model at the beginning of the RH ramp in Figs. 7 and 8, respectively.

Perhaps a more complex model, which divides the body undergoing evaporative cooling

into a shell and a core (van der Sman, 2003), would have reduced the initial discrepancy.

The average absolute errors between prediction and measurement in Figs. 7 and 8 are 0.3

°C and 0.1–0.3 °C for the van der Sman and simple models, respectively.

The efficacy of evaporative cooling under tropical conditions with a prototypical

ECC (Fig. 3) can be observed from the data presented in Table 5. The data, which were

20
taken in Kolkata, India on the afternoon of May 13, 2004 when the ambient temperature

in the room was 34.5–35 °C, show that as the afternoon progressed, the cooling effect

(Tamb – Twater) reached a maximum of 7 °C between 3 and 3:30 PM (mid afternoon, RHamb

= 63%) and subsequently declined to 4.5 °C around 6 PM (evening, RHamb = 88%). The

cooling effect would probably have been larger if the experiment had been conducted

outdoors. The cooling effect of 7 °C at RHamb = 63% observed under tropical conditions

(Tamb = 35 °C) can be compared to the one of 4.7 °C at RHamb = 60% observed under

more temperate conditions (Tamb = 23.4 °C; see Table 4).

SUMMARY AND CONCLUSIONS

In this study, steady-state and dynamic experiments were performed to investigate

the evaporative cooling of water contained in a small porous clay vessel kept inside a

controlled humidity chamber whose RH was either maintained at a constant value or

varied linearly with time from one prescribed level to another. In the steady-state

experiments as the RH of the chamber, whose temperature was 23 °C, was decreased

from 60 to 15%, the cooling effect due to evaporation, represented by the difference

between the ambient temperature and the temperature of the water in the clay pot,

increased from 4.7 to 8.3 °C, which shows the critical influence of the ambient RH on the

temperature depression. The steady-state data were used to estimate the heat and mass

transfer coefficients outside the vessel, which were subsequently employed in a

mathematical model recently proposed in the literature (van der Sman, 2003) to predict

the transient temperature variation of the water in the vessel as the ambient RH was

reduced from 60 to 15% in a linear fashion with time. A very simple mathematical model

was also developed to describe the temperature evolution of an object cooled by

21
convection and evaporation at its surface under a dynamically changing ambient RH. In

contrast to the model of van der Sman, this model requires the ratio of the mass and heat

transfer coefficients, contains an empirical fitting parameter (hint) but has a closed form

solution. Except for an initial discrepancy in van der Sman’s model, on average, both

dynamic models gave predictions that were in fair agreement with experiment.

Evaporative cooling under tropical conditions was investigated with a prototypical ECC

in Kolkata, India. During an afternoon of May 2004 when the ambient temperature was

34.5–35 °C, the cooling effect reached a maximum of 7 °C between 3 and 3:30 PM and

then declined to 4.5 °C around 6 PM.

This work confirms previous findings in the literature that a cooling effect of

several degrees Celsius can be realized through evaporative cooling and this effect will

be enhanced in a drier environment. It is hoped that the two models of the evaporative

cooling process examined in this work will be useful for the rational design of

evaporatively cooled chambers and hollow fibre membrane or porous ceramic units for

room air conditioning and for estimating the cooling time of food products.

NOMENCLATURE

a Parameter in Eq. (28) (K or °C)

Aext External surface area of body (cm2)

ap Half thickness or radius of body (cm)

b Parameter in Eq. (28) (K/cmn or °C/cmn)

camb Ambient water vapor concentration (gmol/cm3)

cp,clay Heat capacity of clay [J/(g K)]

cp,eff Effective heat capacity of body [J/(g K)]

22
cp,water Heat capacity of water [J/(g K)]

cs Water vapor concentration at the surface of body (gmol/cm3)

c1, c2 Constants in Eq. (10) (atm, K-1)

c3, c4 Constants in Eq. (11) [J/g, J/(g K)]

c5, c6 Constants in Eq. (24) [cm/(s %), cm/s]

c7, c8 Constants in Eq. (25) [cm3 K/(J %), cm3 K/J]

c9, c10 Constants in Eq. (34) (°C/%, °C)

c11, c12 Constants defined by Eqs. (36) and (37), respectively (s-1, K or °C)

dc Characteristic length (cm)

f Function defined by Eq. (7)

g Function defined by Eq. (9)

hext External heat transfer coefficient [W/(cm2 K)]

hf Height of frustum of the clay pot (cm)

hi Effective internal heat transfer coefficient defined by Eq. (16) (s-1)

hint Internal heat transfer coefficient defined by Eq. (5) [W/(cm2 K)]

kclay Thermal conductivity of fire clay [W/(cm K)]

keff Effective thermal conductivity of body [W/(cm K)]

kwater Thermal conductivity of water [W/(cm K)]

kf,ext External mass transfer coefficient (cm/s)

lcyl Exposed length of upper cylindrical portion of clay pot (cm)

mclay Mass of clay pot (g)

mwater Average mass of water in clay pot (g)

Mair Average molecular weight of air (29 g/gmol)

23
Mw Molecular weight of water (18.016 g/gmol)

n Constant (≥ 1) in Eq. (28)

Psat(T) Vapor pressure of water at temperature T (atm)

r Ramp rate (s-1)

rcyl Outer radius of upper cylindrical portion of clay pot (cm)

rbot, rtop Bottom and top radii of frustum of clay pot, respectively (cm)

Requiv Equivalent radius of body (cm)

Rgas Universal gas constant [82.0562 cm3 atm/(gmol K)]

Rwater Defined by Eq. (22) (cm)

RHamb Ambient relative humidity (%)

RHi, RHf Initial and final relative humidities at the beginning and the end of the RH

ramp, respectively (%)

t Time (s)

tr Duration of RH ramp given by Eq. (26) (s)

twall Wall thickness of clay pot (cm)

T Temperature (K)

Tamb Ambient temperature (K or °C)

Tavg Average temperature of body (K or °C)

Tavg ,tr Average temperature of body at t = tr (K)

Tcntr Temperature on the central axis of symmetry of body (K or °C)

Teq,f Equilibrium temperature of body corresponding to RHf (K)

Teq Equilibrium temperature of body (K or °C)

Tini Initial temperature of body (K)

24
Ts Surface temperature of body (K)

Tw Local temperature inside body at radial position x and time t (K or °C)

Twater Temperature of water near the top and center of the prototypical ECC (°C)

V Volume of body (cm3)

Wloss Weight-loss rate of body due to evaporation under steady-state conditions (g/s or

g/h)

x Radial distance measured from axis of symmetry of body (cm)

αeff Effective thermal diffusivity of body [keff/(ρeff cp,eff), cm2/s]

β Time constant defined by Eq. (33) (s-1)

∆T Temperature depression, Tamb – Teq (K or °C)

φext Net external heat flux at the surface of body (J/s)

λ Latent heat of vaporization of water (J/g)

ρair,amb Density of ambient air (g/cm3)

ρeff Effective density of body (g/cm3)

REFERENCES

Aimiuwu, V. O. 1992. Evaporative cooling of water in hot arid regions. Energy

Conversion and Management 33(1): 69-74.

Aimiuwu, V. O. 1993. Ceramic storage system based on evaporative cooling. Energy

Conversion and Management 34(8): 707-710.

Anyanwu, E. E. 2004. Design and measured performance of a porous evaporative cooler

for preservation of fruits and vegetables. Energy Conversion and Management 45:

2187-2195.

25
Armbruster, R. and J. Mitrovic. 1998. Evaporative cooling of a falling water film on

horizontal tubes. Experimental Thermal and Fluid Science 18: 183-194.

Bergero, S. and A Chiari. 2001. Experimental and theoretical analysis of air

humidification/dehumidification processes using hydrophobic capillary

contactors. Applied Thermal Engineering 21: 1119-1135.

Chatterjee, S. G., B. V. Ramarao and C. Tien. 1997. Water-vapour sorption equilibria of

a bleached-kraft paperboard – a study of the hysteresis region. Journal of Pulp

and Paper Science 23(8): 366-373.

Cheikh, H. B. and A. Bouchair. 2004. Passive cooling by evapo-reflective roof for hot

dry climates. Renewable Energy 29: 1877-1886.

Chuntranuluck, S., C. M. Wells and A. C. Cleland. 1998a. Prediction of chilling times of

foods in situations where evaporative cooling is significant − Part 1. Method

Development. Journal of Food Engineering 37: 111-125.

Chuntranuluck, S., C. M. Wells and A. C. Cleland. 1998b. Prediction of chilling times of

foods in situations where evaporative cooling is significant − Part 2. Experimental

Testing. Journal of Food Engineering 37: 127-141.

Chuntranuluck, S., C. M. Wells and A. C. Cleland. 1998c. Prediction of chilling times of

foods in situations where evaporative cooling is significant − Part 3. Applications.

Journal of Food Engineering 37: 143-157.

Dai, Y. J. and K. Sumathy. 2002. Theoretical study on a cross-flow direct evaporative

cooler using honeycomb paper as packing material. Applied Thermal Engineering

22: 1417-1430.

26
Dash, S. K. and P. Chandra. 2001. Effects of different structural and operational

parameters on the thermal environment of an evaporatively cooled storage

structure. International Agricultural Engineering Journal 10(3-4): 231-253.

Dhemre, J. K. and D. P. Waskar. 2003. Effect of post-harvest treatments on shelf-life and

quality of mango in evaporative cool chamber and ambient conditions. Journal of

Food Science and Technology 40(3): 316-318.

Dzivama, A. U., F. O. Aboaba and U. B. Bindir. 1999. Evaluation of pad materials in

construction of active evaporative cooler for storage of fruits and vegetables in

arid environments. Agricultural Mechanization in Asia, Africa and Latin America

30(3): 51-55.

Fu, W-S., J-D. Lin, K-C. Tu, and C-C. Tseng.1990. Numerical study of a structure

protected by water film from an incident heat flux. International Communications

in Heat and Mass Transfer 17(3): 331-342.

Ghiabaklou, Z. 2003. Thermal comfort prediction for a new passive cooling system.

Building and Environment 38: 883-891.

Giabaklou, Z. and J. A. Ballinger. 1996. A passive evaporative cooling system by natural

ventilation. Building and Environment 31(6): 503-507.

Ibrahim, E., L. Shao and S. B. Riffat. 2003. Performance of porous ceramic evaporators

for building cooling application. Energy and Buildings 35: 941-949.

Johnson, D. W., C. Yavuzturk and J. Pruis. 2003. Analysis of heat and mass transfer

phenomena in hollow fiber membranes used for evaporative cooling. Journal of

Membrane Science 227: 159-171.

27
Kassem, A-W. S. 1994. Energy and water management in evaporative cooling systems in

Saudi Arabia. Resources, Conservation and Recycling 12: 135-146.

Kumar, A., B. S. Ghuman and A. K. Gupta. 1999. Non-refrigerated storage of tomatoes –

effect of HDPE film wrapping. Journal of Food Science and Technology 36(5):

438-440.

Liu, W., S. W. Peng and K. Mizukami. 1997. Moisture evaporation and migration in thin

porous packed bed influenced by ambient and operating conditions. International

Journal of Energy Research 21: 41-53.

Mordi, J. I. and A. O. Olorunda. 2003. Effect of evaporative cooler environment on the

visual qualities and storage life of fresh tomatoes. Journal of Food Science and

Technology 40(6): 587-591.

Nahar, N. M., P. Sharma and M. M. Purohit. 1999. Studies on solar passive cooling

techniques for arid areas. Energy Conversion & Management 40: 89-95.

Perry, R. H., D. W. Green and J. O. Maloney, Eds. 1984. Perry’s Chemical Engineers’

Handbook. Sixth edition. New York: McGraw-Hill.

Prange, H. D. 1996. Evaporative cooling in insects. J. Insect Physiol. 42(5): 493-499.

Song, C. H., D-Y. Lee and S. T. Ro. 2003. Cooling enhancement in an air-cooled finned

heat exchanger by thin water film evaporation. International Journal of Heat and

Mass Transfer 46: 1241-1249.

Taha, A. Z., A. A. A. Rahim and O. M. M. Eltom. 1994. Evaporative cooler using a

porous material to be used for reservation of food. Renewable Energy 5(1): 474-

476.

28
Tang, R. and Y. Etzion. 2004. On thermal performance of an improved roof pond for

cooling buildings. Building and Environment 39: 201-209.

Thakur, K. S., B. B. Lal Kaushal and R. M. Sharma. 2002. Effect of different post-

harvest treatments and storage conditions on the fruit quality of kinnow. Journal

of Food Science and Technology 39(6): 609-618.

Tien, C.1994. Adsorption Calculations and Modeling. Boston: Butterworth-Heinemann.

Treybal, R. E. 1980. Mass-Transfer Operations. New York: McGraw-Hill.

Upchurch, D. R. and J. R. Mahan. 1988. Maintenance of constant leaf temperature by

plants − II. Experimental observations in cotton. Environmental and Experimental

Botany 28(4): 359-366.

Uppal, D. S. 1999. Effect of storage environments on chip colour and sugar levels in

tubers of potato cultivars. Journal of Food Science and Technology 36(6): 545-

547.

van der Sman, R. G. M. 2003. Simple model for estimating heat and mass transfer in

regular-shaped foods. Journal of Food Engineering 60: 383-390.

Waskar, D. P., R. M. Khedkar and V. K. Garande. 1999. Effect of post-harvest treatments

on shelf life and quality of pomegranate in evaporative cool chamber and ambient

conditions. Journal of Food Science and Technology 36(2): 114-117.

Zalewski, W. and P. A. Gryglaszewski. 1997. Mathematical model of heat and mass

transfer processes in evaporative fluid coolers. Chemical Engineering and

Processing 36: 271-280.

29
Table 1. Characteristic length dc for regular-shaped objects in the linear driving force

approximation.

Geometry Characteristic length dc

van der Sman (2003) Glueckauf

Slab (double sided cooling) ap/2 ap/3

Cylinder ap/3 ap/4

Sphere ap/4 ap/5 (Tien, 1994)

Table 2. Physical properties and relationships.

Property or parameter Relation Values

Vapor pressure of water, c1 = 2.607 × 10 -10 atm


Psat (atm)a, Eq. (10) c1ec2T c2 = 0.0624 K-1
(valid for T = 287–297 K)
Latent heat of vaporization c3 = 3151.5 J/g
of water, λ (J/g)a, Eq. (11) c3 + c4T c4 = -2.38 J/(g K)
(valid for T = 285–295 K)
Heat capacity of water, 4.187 J/(g K)
cp,water [J/(g K)]
Thermal conductivity of 6.15 × 10-3 W/(cm K)
water, kwater [W/(cm K)]
External mass transfer c5 = 0.014 cm/(s %)
coefficient, kf,ext (cm/s), c5 RH amb + c6 c6 = 0.203 cm/s
Eq. (24) (valid for RHamb = 15–60%
in this work)
Ratio of external mass and c7 = 4.941 cm3 K/(J %)
heat transfer coefficients, c7 RH amb + c8 c8 = 264.517 cm3 K/J
kf,ext/hext (cm3 K/J), Eq. (25) (valid for RHamb = 15–60%
in this work)
Wet-bulb or equilibrium c9 = 0.085 °C/%
temperature, Teq (°C), c9 RH amb + c10 c10 = 13.369 °C
Eq. (34) (valid for RHamb = 15–60%
in this work)
a
Parameters were obtained by regression of values given by Perry et al. (1984).

30
Table 3. Parameters of the clay pot shown in Fig. 1.

Property Type or Value

Material Porous red clay

Exposed length of upper cylindrical portion, lcyl 1.1 cm

Outer radius of upper cylindrical portion, rcyl 2.4 cm

Height of frustum, hf 2.7 cm

Outer radius of frustum at top, rtop 2.2 cm

Outer radius of frustum at bottom, rbot 1.5 cm

Wall thickness, twall 0.3 cm

External surface area, Aext [from Eq. (19)] 51.9 cm2

Mass of clay pot, mclay 22.681 g

Heat capacity of clay, cp,claya 0.938 J/(g K)

Thermal conductivity of (fire) clay, kclaya 10.02 × 10-3 W/(cm K)


a
From Perry et al. (1984).

31
Table 4. Results of the steady-state evaporative cooling experiments with the clay pot

containing water.

RHamb (%) Tamb (°°C) Teq (°°C) Wloss (g/h)

Replicate #1

59.8 23.3 18.7 0.652

45.0 23.2 17.0 0.744

30.1 23.1 15.9 0.829

15.0 23.1 14.8 0.790

Replicate #2

59.8 23.4 18.7 0.708

45.0 23.0 16.7 0.791

30.0 23.0 16.0 0.795

15.0 22.8 14.7 0.789

Average

59.8 23.4 18.7 0.680

45.0 23.1 16.8 0.757

30.0 23.0 15.9 0.812

15.0 23.0 14.7 0.790

32
Table 5. Observed cooling effect under tropical conditions with the prototypical ECC
shown in Fig. 3.

Timea Tamb (°°C) RHamb (%) Twater (°°C) Tamb – Twater (°°C)

2:30 PM 34.5 63.5 28.5 6


3:00 PM 34.5 62.5 27.5 7
3:30 PM 35 63 28 7
4:00 PM 34.5 62 28 6.5
4:30 PM 35 74.5 29 6
5:00 PM 35 86 30 5
5:30 PM 34.5 87 30 4.5
6:07 PM 34.5 88 30 4.5
a
The experiment was started at 2:03 PM.

33
List of Figures

Figure 1. (a) Photograph of the small clay pot used in the evaporative cooling

experiments, and (b) schematic diagram of the clay pot (dimensions are given in Table 3).

Figure 2. Schematic diagram of the controlled humidity chamber in which the

evaporative cooling experiments with the clay pot containing water were performed.

Figure 3. Schematic diagram of the prototypical ECC tested under tropical conditions.

Figure 4. Variation of the ratio of the external mass and heat transfer coefficients of the

clay pot containing water with ambient relative humidity under steady-state conditions

(Tamb = 22.8–23.4 °C).

Figure 5. Variation of the external mass transfer coefficient of the clay pot containing

water with ambient relative humidity under steady-state conditions (Tamb = 22.8–23.4 °C).

Figure 6. Variation of the steady-state equilibrium temperature of the clay pot containing

water with ambient relative humidity (Tamb = 22.8–23.4 °C).

Figure 7. Transient central temperature of the clay pot containing water as a function of

time at a ramp rate of 0.5% RH/min [RHi = 60%, RHf = 15%, Tamb = 22.3 °C, mwater =

(23.369 g + 21.848 g)/2 = 22.609 g]. For the simple model, hint = 1.817 × 10-3 W/(cm2 K).

Figure 8. Transient central temperature of the clay pot containing water as a function of

time at a ramp rate of 1% RH/min [RHi = 60%, RHf = 15%, Tamb = 22.2 °C, mwater =

(25.239 g + 24.009 g)/2 = 24.624 g]. For the simple model, hint = 1.817 × 10-3 W/(cm2 K).

34
(a)

Styrofoam lid
rcyl

rtop lcyl

hf

rbot

(b)

Figure 1. (a) Photograph of the small clay pot used in the evaporative cooling
experiments, and (b) schematic diagram of the clay pot (dimensions are given in Table
3).

Electronic
balance

35
Computer Compressed air
Balance Dry
air Dryer
DAV
Motor
H
Humid
air Pump
Trap
WAV
S C S
P Room air
Humidifier
Mixed air
C Clay pot with water
DAV Dry air valve
H Humidity sensor
P Pan
S Screen
WAV Wet air valve

Figure 2. Schematic diagram of the controlled humidity chamber in which the


evaporative cooling experiments with the clay pot containing water were performed.

36
Copper vessel wrapped with a
piece of wet cloth
Wooden board

Fan Fan

Figure 3. Schematic diagram of the prototypical ECC tested under tropical conditions.

37
580
From expt. data and Eq. (17)
Eq. (25)
530
kf,ext /hext [K cm /J]
3

480
y = 4.941x + 264.517
2
R = 0.960
430

380

330
10 20 30 40 50 60 70
RH amb (%)

Figure 4. Variation of the ratio of the external mass and heat transfer coefficients of
the clay pot containing water with ambient relative humidity under steady-state
conditions (Tamb = 22.8–23.4 °C).

38
1.10
From expt. data and Eq. (18)
1.00
Eq. (24)
0.90
kf,ext (cm/s)

0.80
y = 0.014x + 0.203
0.70 2
R = 0.984
0.60

0.50

0.40
10 20 30 40 50 60 70
RH amb (%)

Figure 5. Variation of the external mass transfer coefficient of the clay pot containing
water with ambient relative humidity under steady-state conditions (Tamb = 22.8–23.4
°C).

39
19
expt. data
Eq. (34)
18

17
Teq (C)

y = 0.085x + 13.369
2
R = 0.977
16

15

14
10 20 30 40 50 60 70
RH amb (%)

Figure 6. Variation of the steady-state equilibrium temperature of the clay pot


containing water with ambient relative humidity (Tamb = 22.8–23.4 °C).

40
20 Expt. data
van der Sman's model
19 Simple model

18
Tcntr (C)

17

16

15

14
0 20 40 60 80 100 120 140 160
t (min)

Figure 7. Transient central temperature of the clay pot containing water as a function
of time at a ramp rate of 0.5% RH/min [RHi = 60%, RHf = 15%, Tamb = 22.3 °C, mwater
= (23.369 g + 21.848 g)/2 = 22.609 g]. For the simple model, hint = 1.817 × 10-3
W/(cm2 K).

41
20 Expt. data
van der Sman's model
19 Simple model

18
Tcntr (C)

17

16

15

14
0 20 40 60 80 100 120 140 160
t (min)

Figure 8. Transient central temperature of the clay pot containing water as a function
of time at a ramp rate of 1% RH/min [RHi = 60%, RHf = 15%, Tamb = 22.2 °C, mwater =
(25.239 g + 24.009 g)/2 = 24.624 g]. For the simple model, hint = 1.817 × 10-3 W/(cm2
K).

42

View publication stats

You might also like